• Aucun résultat trouvé

A maximum entropy principle in general relativity and the stability of fluid spheres

N/A
N/A
Protected

Academic year: 2022

Partager "A maximum entropy principle in general relativity and the stability of fluid spheres"

Copied!
25
0
0

Texte intégral

(1)

A NNALES DE L ’I. H. P., SECTION A

W. J. C OCKE

A maximum entropy principle in general relativity and the stability of fluid spheres

Annales de l’I. H. P., section A, tome 2, n

o

4 (1965), p. 283-306

<http://www.numdam.org/item?id=AIHPA_1965__2_4_283_0>

© Gauthier-Villars, 1965, tous droits réservés.

L’accès aux archives de la revue « Annales de l’I. H. P., section A » implique l’accord avec les conditions générales d’utilisation (http://www.numdam.

org/conditions). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright.

Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques

http://www.numdam.org/

(2)

283

A maximum entropy principle

in general relativity

and the stability of fluid spheres (1)

W.

J. COCKE

(Cornell University (N. Y)

et Institut Henri

Poincaré, Paris)

Ann. Inst. Henri Poincaré, Vol. II, no 4, 1965,

Section A :

Physique théorique.

SUMMARY. - The

principle

that the total

entropy

be a maximum is

applied

to

gravitational systems

of a

perfect,

relativistic fluid in adiabatic motion.

An

expression

for the

entropy density

of the fluid is derived as an

explicit

function of the energy

density,

and the conservation

equations imply

that

this

expression

satisfies the proper

continuity equation.

The

entropy principle

is shown to

yield

the

equation

of

hydrostatic equilibrium

and a

stability

criterion

equivalent

to one derived from

dynamics. Application

to features of

equilibrium

models is studied.

SOMMAIRE. - Le

principe d’entropie

maximum est

applique

au

système gravitationnel :

fluide relativiste

parfait

en mouvement

adiabatique.

On trouve une

expression

de la densite

d’entropie,

fonction

explicite

de la

densite

d’energie.

Les

equations

de conservation entrainent que cette fonc- tion satisfait

identiquement

a une

equation

de continuité. On deduit du

principe d’entropie

maximum

1’equation d’équilibre hydrostatique,

et un

critère de stabilité

equivalent

a celui

qui

découle de la

dynamique.

Des

applications

aux modèles

d’équilibre

sont etudiees.

(1)

This paper is a revised version of a thesis presented to the

Faculty

of the

Graduate School of Cornell

University,

Ithaca, New York, U. S. A., in

partial

fulfillment of the

requirements

for the degree of Doctor of

Philosophy.

The author is

presently

the holder of a Postdoctoral

Fellowship

of the National Science Foundation.

ANN. INST. POINCARÉ, A-II-4 19

(3)

284 W. J. COCKE

INTRODUCTION

This paper deals

chiefly

with

setting

up a maximum entropy

principle

for

gravitationally

bound

spheres

of

perfect fluid,

treated via Einstein’s

general theory

of

relativity.

We will derive an

expression

for the total

entropy

of such a fluid

sphere

in terms of the pressure, energy

density,

and

gravitational potentials

and

require

that this

expression

be a maximum

with

respect

to small

perturbations conserving

the total energy of the fluid.

We will show that this

requirement implies

both the relativistic

equation

of

hydrostatic equilibrium

and a criterion for the

stability

of static

configu-

rations of fluid

satisfying

this

equation.

Considerable interest will lie in the

stability criterion,

and it will be shown that it is

equivalent

to one

previously

obtained

by

S. Chandrasekhar from

dynamical

considerations. We will also use it to elucidate certain features of actual

equilibrium

models. In

particular,

we will consider how unstable

configurations

are related to the location of maxima and minima of

plots

of

equilibrium

mass curves,

parametrized

as functions of the central pressure or

density.

Throughout

the paper we consider all motions to be

adiabatic,

so that

we

neglect

heat

flow,

but we make no other

assumption

as to the form of the

equation

of state,

except

that it be a function of the usual

thermodynamic

variables. We will limit our discussion to

stability against spherically symmetric perturbations, although

the formalism could be extended to include other

types.

In

general

there are two methods of

testing

the

stability

of a static confi-

guration

of

fluid,

both of which have been

applied

to the Newtonian case.

We will now

briefly

discuss these two methods and indicate the connection between them.

The first method is known as the «

dynamical

method » and has been

applied

to both Newtonian and relativistic fluid mechanics

[1, 2].

It

involves

assuming

that the

quantities 03C8j characterizing

the state of the fluid

system deviate,

as functions of

positions

and

time, only

very

slightly

from

.a set of

given

static

quantities

One may then take the

time-dependent equations

of motion to be linear

equations

with the small differences

~ 2014

==

~

as

dependent

variables.

If the motions are assumed to be

adiabatic,

and hence

reversible,

the

time-dependance

can be

separated

out

by assuming

a functional

dependence

on r, t of the form

~y

= cos at, where a is a constant. The

equations

(4)

285

A MAXIMUM ENTROPY PRINCIPLE

for

~~~(r)

can then be transformed into a Sturm-Liouville

eigenvalue

pro- blem, with a2 as

eigenvalue.

See in

particular

section I of this paper, where

we

present

a sketch of the

dynamical

method in

general relativity.

In

general

the admissible

eigenvalues

a2 form a

discrete, countably

infinite

set, and

stability

is tested

by ascertaining

whether or not all the a2’s are

positive.

If not, then some of the a’s are

imaginary,

and excitation of these modes will result in an

exponential time-dependence

instead of an oscil-

latory

one. This will carry the

configuration

so far away from the

given

static state that the assumed linear

approximation

will become invalid.

A static

configuration

is then stable with

respect

to small

perturbations only

if all the

eigenvalues s2

associated with it are

positive.

The second method is known as the « energy method » and has so far

only

been

applied

in Newtonian

theory [3, 4, 5].

It states the

following:

if the total conserved energy E of a mass of fluid can be written as a sum

E = T +

U,

where T is a

positive

definite kinetic energy and U is a

poten-

tial energy which does not

depend

on the

hydrodynamic velocities,

then a

configuration

is in stable

equilibrium only

if U is a minimum with

respect

to infinitesimal variations of the

quantities ~j

which conserve the total

amount

of

material

(in

Newtonian

theory,

the total

mass).

That this is

necessary for the

stability

of a static distribution of material is

evident,

since otherwise any

lessening

of the

potential

energy U

by

a small

pertur-

bation would release kinetic energy and

provide

further

impetus

to the

displacement.

In terms of the calculus of

variations, requiring

U to be a minimum means

that we must have 3D = 0 and ö2U &#x3E;

0,

where ~U is the first variation and ö2U is the second variation. au = 0

implies

an Euler

equation,

which

turns out to be

simply

the

hydrostatic equilibrium equation,

and a2D &#x3E; 0 is then a

stability

criterion.

The connection between the

dynamical

method and the energy method is made as follows : it turns out

[6]

that a necessary and sufficient condition for 82U &#x3E; 0 for non-zero variations is that the

eigenvalues

fL of a

particular eigenvalue problem

all be

positive.

This

eigenvalue problem

is in

general

not

quite

the same as the one associated with the

dynamical

method

(as

discussed

above),

but one should be able to show that all

the , ’s

are

positive

if and

only

if all the a2’s are

positive.

This would show that the two stabi-

lity

criteria are in fact the same. This has been done in the Newtonian

case

[5, 6].

In

general relativity, however,

the formulation of the energy method is

complicated by

the

apparent

lack of a

quantity characterizing

the « total

(5)

286 W. J. COCKE

amount of material ». In Newtonian mechanics this

quantity

is

simply

the total mass of the

fluid,

but in

relativity

the fundamental

equivalence

of mass and energy

implies

that one could alter the total mass of a

body

without

actually removing

or

adding

any « material».

Many

authors

have used the

concept

of a total

particle

number or so-called « proper mass », but this usage seems

foreign

to

hydrodynamics,

which is a

theory

of conti-

nuous fields

[7].

We may,

however,

turn to the

thermodynamic aspect

of the

problem

and

investigate

the

requirement

that the total

entropy

of the

system

be a

maximum for

stability.

It is this

principle which,

as stated

above,

we will

develope

in this paper, and with which we propose to

replace

the minimum

potential

energy method.

Since we consider

only

adiabatic

motions,

our

expression

for the total

entropy

S will turn out to be a constant of the motion. We will derive

an

expression

for the

entropy density

in terms of

quantities entering directly

into Einstein’s field

equations,

and then

verify

from the- field

equations

that it satisfies the proper

continuity equation.

An

expression

for the conser-

ved total

entropy

S will then follow. This will

give

us a non-trivial

integral

of Einstein’s field

equations

for

general

adiabatic fluid motions.

We will next show that the statement 8S =

0,

taken relative to variations which conserve the total energy,

implies

the relativistic

equation

of

hydro-

static

equilibrium.

The additional maximum

requirement

82S 0 will

then be shown to lead to the same

stability

criterion obtained

by

Chan-

drasekhar from the

dynamical

method. In the last

section,

the maximum entropy

principle

will then be

applied

to some other

interesting

ques- tions.

At this

point

we

might

say a few words

regarding

the

assumption

that

no heat flow occurs. It is of course evident that this

postulate

is never

exactly

satisfied in nature, but we may

regard

it as

approximately

true over

appropriately

short intervals of proper time relative to an observer

moving

with the fluid. In situations

involving

very strong

gravitational fields,

such as in a

gravitational collapse problem,

this

approximation

would

appear to be a propos, in as much as proper times inside the fluid would be much slowed down

compared

to the proper times of external observers.

However,

we will not here

investigate

the

validity

of our

assumptions,

but will content ourselves with

exploring

their consequences. It would be

interesting

to see how the methods of this paper

might

be extended to include the flow of heat and radiation

coupled

with the matter fields.

(6)

287

A MAXIMUM ENTROPY PRINCIPLE

We use the

following

notation and conventions. The

expression

for

the interval

length

will be written as

with Greek indices

taking

on the values

0, 1, 2,

3. Latin indices will be understood to assume the values

1, 2, 3.

Thus the fluid

4-velocity

satisfies = 1.

We choose units so that the

velocity

of

light

c = 1 and the Newtonian

gravitational

constant G = 1. Thus we will write the field

equations

as

where is the Einstein tensor, which satisfies the conservation

identity

V

03B1S 03B1

=

0, ~

«

denoting

the

operation

of covariant

differentiation,

and

is the energy-momentum tensor.

The

energy-momentum

tensor of the

perfect

fluid

without

heat flow is

then

expressed

in these units as

where e is the total proper energy

density, p

is the pressure, and is the

hydrodynamic

flow

4-velocity.

. 1. - THE DYNAMICAL METHOD

IN GENERAL RELATIVITY

To

bring

into focus some of the ideas of the

introduction,

we here

present

a sketch of the

dynamical

method of

testing

the

stability

of fluid

spheres

in

general relativity.

We have mentioned that we will

only

consider

stability against spherically symmetrical perturbations,

and since any static

configu-

ration of fluid will exhibit such symmetry, we way

conveniently

take this

as a

starting point.

It will be useful to

employ spherical polar

coordinates and to define the radius

parameter

r

by using

the metric form

v and x

being

real functions of r and t.

Einstein’s field

equations Sf

= - then

give [8], using

a

prime

to

denote

()-/o-r, for l.L

= v =

1,

(7)

288 W. J. COCKE

and

for y = v = 0,

where we have used = 1.

Integration

of

Eq. (1-3) yields

For y

=

1, v

=

0, we

obtain

where the dot indicates The

equation for ,

= v = 2 would

provide

a fourth

equation (There

are four

independent unknowns,

which may be taken to be e,

ul,

x, and v, p

being given

as a function of the other variables.

See section

II). However,

it is more convenient to use

the l-L

= 1

component

of the conservation

equation V

= 0. We shall not write it out in

full,

but rather refer the reader to the literature

[9].

We now denote

where so, po,

a~,

vo are solutions of the static

equations

or

and of the static

part

of the conservation

equation

=

0,

which reads

Equation (1-6)

is the

equation

of

hydrostatic equilibrium,

which we will

derive

again

from the maximum entropy

principle

in section III. It

is,

of course, the

analogue

of the Newtonian

equation dp/dr

= -

pdp/dr,

p

being

the mass

density and p

the

gravitational potential.

Following

Chandrasekhar

[7],

we assume that

8e, 8p, 8À, 8v,

and u1 are

(8)

289

A ENTROPY PRINCIPLE

small

quantities

defined on the interval 0 r

Ro,

where

Ro

is the boun-

dary

radius of the static

distribution ;

and we write to first order

of smallness

Letting ~(r, t )

be the «

Lagrangian displacement

)) of an element of the fluid from an initial

position,

we have to first order

and

integrating Eq. ( 1-5),

we

get, with U4

"’

eVo/2,

From V

a Tla

=

0,

to first

order,

one may obtain the relation

Since this

equation

is linear in the

dependent

variables and contains

only

the second time

derivative,

we may

separate

variables

by assuming

a time-

dependence

of the form cos at, a

being

a constant,

getting

where we now consider

8 ~, 8v, 8p, oe, and ~

to be functions of r

only.

In the next section we show that for adiabatic motions we may

write p

in the

form p

=

p(e, x),

where the variable x indexes the fluid element itself and follows its motion. See also reference

[1].

Then we have in

our case

and we may use this relation to substitute for

ap

in

Eq. (1-7),

and

use

Eq. (1-2") (1-4")

and

( 1-5 ")

to eliminate other

variables, finally writing

Eq. (1-7) with ç

as

dependant

variable

(9)

290 W. J. COCKE

where one can check to see that P ~ 0.

Multiplication by

the

quantity q(r)

= exp allows us to write the form

where

q(r), v(r)

&#x3E; 0. The

boundary

conditions

on ç

are

arranged

so that

the pressure vanishes to first order on the new

boundary

radius

Ro

+

ç(Ro).

This condition reads

which the reader may

easily

write himself in terms

of ç.

At r = 0 we must have of

course ç

= 0.

Equations (1-8)

and

(1-9)

constitute a Sturm-Liouville

eigenvalue problem

The

problem

is of the

singular

type

since q

and v may be shown to tend to 0 at both ends of the interval 0 r

Ro. However,

one may show

[2]

that in

spite

of the

singular

character of the

problem,

there exists an infinite discrete set of

eigenvalues { 6~ },

of which there is a least member

ao-

It is

easily

shown

[2, 10]

that we may solve

Eq. (1-8)

for a2 as

thus

defining

the functionals

3)[~], JC[~].

The

boundary

terms vanish

because

q(0)

=

q(Ro)

= 0.

According

to theorems on the extremum pro-

perties

of

eigenvalues [2, 10]

the lowest

eigenvalue 6o

is

given by

where 03C6

ranges over all functions with

bounded, piece-wise

continuous

p’

which

satisfy

the

boundary

conditions

for ç

and do not vanish

identically.

The condition

for

stability

was first

applied

to this

problem by

S. Chandrasekhar

[1].

Note that

since q, v &#x3E; 0,

if t ~ 0 for all 0 ~ r

Ro,

then

G~

&#x3E; 0 would follow

immediately,

and any such

configuration

would be

dynamically

stable.

Having

studied the

application

of the

dynamical

method in

general

rela-

tivity,

we now

proceed

to

develope

our

replacement

of the energy

method,

the maximum entropy

principle.

(10)

291

A ENTROPY PRINCIPLE

2. - THE TOTAL ENTROPY

In the introduction we stated the need for

expressing

the total

entropy

S of the fluid

sphere

in terms of variables

already entering

into the field equa- tions. To

accomplish

this

end,

we now construct a relation between the proper entropy

density s

per unit volume and the pressure p and energy

density

e.

First let V be the volume of a very small element of the fluid. Since we

consider all motions to be

adiabatic,

the total

entropy

S = sV of the ele- ment will be constant over such

motions,

and we may write the

thermodyna-

mic

identity

for the element as dE +

pdV

= TdS

= 0;

or, since E =

sV,

Now since a

change

in volume is the

only

means of

changing

the thermo-

dynamic

state of the fluid

element,

heat flow

having

been

outlawed,

it

follows that any

change

in state will be

accompanied by

a non-zero

change

in e,

except

for elements in

which p

and e are both zero

simultaneously.

Thus, except

for such «

empty » elements,

we may say all the

thermodynamic

variables may be written as functions of the energy

density

e,

including

of course the pressure.

However,

the

equation

of state p =

f(e)

will then in

general

be a different function of e for different fluid elements.

Let us now consider the use of a

co-moving

coordinate

system

in which the

spatial components

uj

(~’ = 1, 2, 3)

of the fluid

4-velocity

vanish

throughout

the

system

for all time. Such coordinates are

presumably always possible

since

they

may be defined

by simply attaching

the

spatial

reference

system

to the identifiable

points

of the fluid.

By

virtue of the above arguments we may now write the pressure as a function of e and of the

spatial

coordinates xj of the

co-moving

frame:

p =

p(e,

the xj

serving

to index the different

points

of the

fluid ;

i. e., the different fluid elements. It must be undertood that the function

p(e, xi)

is to be derived from

purely thermodynamic

considerations after

specifying

the initial pressure and energy

density

distributions at a time x° =

But since the total entropy of the fluid element S = sV is

conserved,

we

may form d S =

d(sV)

= sdV + Vds =

0,

or -

dV/V

=

ds/s. Hence,

from

above,

(11)

292 W. J. COCKE

Integration

with xi = constant

yields

or

where so and so are functions of xj such that

s(eo)

= so.

We have thus

expressed

the

entropy density s

as a definite function of the energy

density e

and the

spatial

co ordinates xj of the

co-moving

frame jc~.

Let it be

emphasized again

that this

expression

is valid

only

for adiabatic

changes

of state.

Note that of it is

possible

to define a conserved

particle

number for the

system such that n is the proper

particle

number

density,

the total

particle

number N = nV of a small fluid element would also be

conserved;

and

likewise, - d V/V

=

dn/n.

Hence we would

similarly

have

°

and thus n and s would be

proportional

to the same function of s for fixed xj.

As an

example

let us find an

expression

for the

entropy density

of a per- fect non-relativistic Boltzmann gas with constant

specific

heat. We may write the

equation

of state for an element of such a gas as

pV

=

NkT,

or p = nkT. For adiabatic

compression

or

expansion

we further

have,

with y as the constant ratio of

specific

heats per

particle

y =

cp/cv [11], pVY

= constant or p = constant nY, or n = where no and pa

are functions of xj. Then the energy

density

is

given by

where mo is the rest mass per

particle

and Eo is an

arbitrary

constant. There-

fore

which definies p =

p(e, xj)

as an

implicit

function. Then

where we have used cp - Cy = k. It then follows that

We have thus verified the

proportionality

of sand n.

(12)

293

A MAXIMUM ENTROPY PRINCIPLE

The total

entropy

of the fluid

element

may be written

[11]

where ~

is the chemical constant of the gas. Hence

This verifies the

expression

for s derived above from

Eq. (2-1).

We must now show that our

expression

for s satisfies the proper conser- vation

equation

= 0. Let us write the scalar in terms of

our

co-moving

frame in which

only

is

non-vanishing :

where indicates Now we have from

Eq. (2-1)

that

so that

and hence

But this is a tensor

equation,

and hence is true not

only

in the

co-moving

frame but in any

system

of coordinates whatever. Now the field

equations imply

that

Let us further construct the scalar

product

um v =

0, using

the rela-

tion = 1 and its

consequent

= 0 :

Comparison

with

Eq. (2-3)

then shows that in

fact,

if the field

equations

are = 0.

(13)

294 W. J. COCKE

We may use this relation to show that the total proper

entropy

of the fluid

system

is a conserved

quantity.

Let t be a time-like coordinate and x, y, z be

space-like

coordinates

(not necessarily

Galilean at

infinity).

Let

2p

be the

space-like

surface of the mass of

fluid,

and let

~a

and

~b

be the intersections of the time-like surfaces t = a and t

= b, respectively,

with

the

space-time

volume of the fluid. Then if

F

be the

part

of

03A3F

which lies between

~a

and

~b,

the surfaces

~Q, ~b,

and

2p together

form a closed

3-space E

in

space-time.

Now let V denote the

space-time

volume of fluid enclosed

by E.

Then

for an

arbitrary

vector field

A~,

we may write Green’s theorem for this volume and surface as

[12]

where n~ is the outward unit normal to

~,

and and are

generalized

3- and 4-volume elements.

If

= 0 is the

equation

of

I:,

then we have

with the

sign

chosen so that n~ is directed outward from E. Inside V

we can let =

il’ -

g

dt dx dy dz

and on

Ea

and

~b

we can set

=

Ý -

g~3~ dx

dy dz,

where is the determinant of the subtensor gij

(i~ .I

=

1, 2, 3).

Then if we choose A~ = we will have = = 0 inside

V,

and

consequently, by

Green’s

theorem,

Now the

equation

for

~a is

= t - a =

0,

so that on

E.,

for a

b,

Similarly,

on

1~

Thus it follows

that,

since

(14)

295

A ENTROPY PRINCIPLE

Further,

we must also

satisfy

the Lichnerowicz

junction

conditions

[13],

from which one may

easily

deduce that on the surface

2p,

ua.na. = 0. This

can also be shown

by expressing

u~ and ntL in

co-moving coordinates,

in

which the

equation

for

Ep

will have the

form f(xj)

= 0.

Thus we obtain

and hence

for all t = a, b.

Or,

where

2~

is the

3-space t

= constant. The

integral

S may be said to repre- sent the total

entropy

in the same sense that conservation of

charge

in elec-

trodynamics

in the form =

0,

where p is the

charge density,

leads to

an

exactly

similar

expression

for the total conserved

charge.

We now

proceed

to formulate our variational

principle

S = maximum

with

respect

to variations which conserve the total energy.

3. - THE FIRST VARIATION

We now

require

that for

hydrostatic equilibrium

the total

entropy

S be an extremum

(maximum)

for all infinitesimal adiabatic variations of the energy

density e

and pressure p which conserve the total energy of the sys- tem. As

stated,

we will restrict ourselves to variations which preserve the

spherical symmetry

of the

fluid,

and thus we will be able to use the expres- sions for the metric form and for the field

equations presented

in section I.

It should be stated at the outset that in

performing

the variations we

will not consider the variations as

independent

from ~e and

8p,

but

will rather let them be

given

via the field

equations (1-2") ( 1-4")

and

(1-5").

We shall not use the

equations T2

= -

803C0S22

and =

0,

since

they

themselves

imply

the

equation

of

hydrostatic equilibrium. Further,

the

variation

öp

will be linked to ~s

by

means of the adiabatic

equation

of state

p =

p( e:, representing,

as it were, the

point

of

origin

of the fluid element. Thus we will have

only

one

independent

variational function.

(15)

296 W. J. COCKE

Let us also note that even

though

the variations are

adiabatic, they

will

not conserve the total

entropy identically,

since we will

always

consider the

spatial velocity

of the material to be zero.

In order to carry out our variations we will need an

expression

for the

total energy. It is

easily

shown that for the metric from

(1-1),

a conserved

energy may be written in the forms

[14]

Thus for the static

problem

we take ui = 0 and write

Using Eq. (2-4),

one may then express the total

entropy,

after

performing elementary integrations,

as

But with u1 =

dr/ds

=

0,

we have from

Eq. (1-1)

that

dt/ds

=

Hence

However,

in

taking

the variations of M and

S,

we must allow for the fact that the

boundary

radius R will also vary. Thus it behooves us to trans- form the

independent

variable r so that the condition M = constant auto-

matically

fixes the end

points

of the interval of

integration.

We do this

by transforming

to the variable

which

represents

the

quantity

of energy inside the radius r.

We denote

dr/dm by

r

(we

trust there will be no confusion with a time

.

derivative in this

context)

and find

dm/dr

= or = r =

and thus

Also we have

(16)

297

A ENTROPY PRINCIPLE

We may hence express the

integrand

of S as a function of the

independent

variable m and the

dependent

variable r and its derivative r. The

princi- ple

~S = 0 will then determine r =

r(m)

from the Euler

equation,

and the

condition M = constant will be

equivalent

to

fixing

the upper end

point

of the interval of

integration

0 m M.

However,

careful note must be made of the fact that the

dependence

of

the

entropy density s

=

xi)

on the

spatial

variables xi is such that xi

always

follows the motion of the fluid element. Thus in

reality,

we have

for the

spherically symmetric problem s

=

s(e,

r -

Q, where ~

is the dis-

placement

of the element of fluid which finds itself at r after

having

been

moved from the

point

r -

ç. ç

is then

given

in terms of the other varia- bles

through

the relation

( 1-5 ").

Let us denote ~ == r -

ç,

so that

s =

x).

Our next

step

would be to write the variation of S in terms of the varia- tion or = of the

dependent

variable r.

However,

we must first be

able to

express ~

in terms of Sr. Here is the infinitesimal

change

of

the radius r =

r(m)

of a

sphere

of fixed energy content m, and not the

Lagrangian displacement,

which we have

designated

as

~.

Let us denote the

change

in energy content

m(r)

of a

sphere

of radius r

by

=

403C0r0

03B4~r2dr. Then

writing mo(r)

as the

equilibrium

distribution

0

function around which the variation is

taken,

we have

by

definition

and hence

to first order.

Thus, using Eq. ( 1-4")

and

( 1-5 "),

we obtain

We now write

(17)

298 W. J. COCKE

The first variation of S will thus be

expressed

as

According

to the fundamental lemma of the calculus of variations the condition 3S = 0

implies

the Euler

equation

and the condition = 0 since in

general ~r(M) 5~ 0, although

of course

8r(0)

= 0.

We first

investigate

the

boundary

term : We have

and

using Eq. (3-3),

which

gives

= -

(403C0r2r2)-1

== -

~/r

one obtains

easily Lr - e03BB/2r2sp(p

+

e)-I.

Thus if the energy

density

does not vanish at the

surface,

we will have

= 0

through

the condition that the pressure on the surface vanish.

But suppose e = 0 also. Then we may write

Lr -

and

demand

simply

that remain finite as e - 0. This is

certainly

a very reasonable

stipulation,

and one can

show,

for

example,

that for

equations

of state of the form e = «p ~--

(see Eq. (2-2)), stays

finite as e2014~0 for

positive

«,

p,

y if and

only

if y &#x3E; 1.

However,

one may also show that the

product

goes to zero as e --~ 0 for all y &#x3E;

0,

so that

Lr]M

= 0

still

tells us very little about the

equation

of state.

We now

proceed

to examine the Euler

equation (3-6).

Let us write

d(Lr )/drn

=

(dr/dm)d(Lr )/dr

= and express

Eq. (3-6)

as

a total differential

equation

with r as

independent

variable as

before, again using

a

prime

to denote

d/dr.

We have

where

Eq. (3-3)

and

(3-4) give

(18)

299

A ENTROPY PRINCIPLE

and

Further

But

so that the terms in will cancel. Then after

using Eq. (1-3’)

and

collecting

terms, we may write

Eq. (3-6)

as

which

with the aid of

Eq. (1-2’)

may be written

p’(p

+ +

v’/2

=

0,

which we

recognize

as

Eq. (1-6),

the

equation

of

hydrostatic equilibrium.

Thus we have obtained

Eq. (1-6) by

means of the maximum

entropy principle

and with the

help

of the

equations S~

= -

8?rT~ S~

= -

and

S~

= - We now

proceed

to examine the second variation ~2S and

require

that it be

negative.

4. - THE SECOND VARIATION AND

EQUIVALENCE

WITH THE DYNAMICAL METHOD

Now that we have found a necessary condition

(the

Euler

equation)

for S

to be an extremum, the

principle

S = maximum

requires

us to

investigate

82S 0 for infinitesimal variations about the extremum. We will show that this

requirement

is

equivalent

to be criterion

developed by

Chandra-

sekhar

by

means of the

dynamical

method. It will be shown in fact that if

~o

be the least

eigenvalue satisfying Eq. (1-8),

then

ao

&#x3E; 0 if and

only

if a2s 0 for all non-zero infinitesimal variations

conserving

total energy and

satisfying

the usual

continuity

and

boundary

conditions.

We will use

again

the

independent

variable m as

developped

in the last section to accommodate variations in the

boundary

radius R and to

comply automatically

with M = constant

by fixing

the upper end

point

of the inter- val of

integration.

ANN. INST. POINCARÉ. A-II-4 20

(19)

300 W. J. COCKE

The second variation 82S is

given by [6, 16]

Substitution of ~r

for ~

via

Eq. (3-5)

then allows us to write 82S in the form

thus

defining

the

quadratic

form

Q(8r, 03B4r).

A rather

simple

test to be satisfied for S to be a maximum is the so-called

Legendre-Weierstrass

test, which is a necessary condition for 828 0 with

respect

to both

strong

and weak variations. It is a local test and is writ- ten

[16, 17]

where rand

r

are actual extremal

values, p

is any

physically

realizable value of r, and 6 is any number between zero and one. The

equality sign

can

hold

only

for isolated

points.

But

where the

dependence

on

r

is

through

£ =

r),

and hence for

positive

£

the test is satisfied if &#x3E;

0,

which

certainly

holds for any real one-

phase

fluid. This also shows that S

certainly

cannot be a minimum.

We now

proceed

to formulate a more

general

test

(related

to the so-called Jacobi test

[18])

which includes the

Legendre-

Weierstrass test in its formula- tion and which we will show is

equivalent

to Chandrasekhar’s criterion.

First we demonstrate that a necessary and sufficient condition for ö2S 0 is that all the

eigenvalues ~

of the associated linear

problem (abbreviating

etc.)

must be

positive.

The

boundary

conditions to be

applied

in

finding

the are

given

as

the latter of which may be shown to be the same as

Eq. (1-9),

which states

that the pressure must vanish to first order on the

varying boundary

of the

sphere.

(20)

301

A MAXIMUM ENTROPY PRINCIPLE

Now let us

multiply Eq. (4-1) by

or and

integrate, getting

Again exploiting

the extremal

properties

of

eigenvalues,

one may

easily

show that for the smallest

eigenvalue

ll-o,

for all

admissible 03C6

not

identically

zero

satisfying

the same

boundary

and

continuity

conditions as ~r. Thus if yo &#x3E;

0,

then 82S

0,

and

conversely.

We are now in a

position

to compare our condition 82S 0 with Chan- drasekhar’s

criterion,

which as we have said results from

linearizing

the

time-dependent equations

of motion.

Substituting ~

for Sr in

Eq. (4-1)

with the use of

Eq. (3-5),

one may, after

changing

the

independent

variable

back to r and

performing

very tedious

manipulations,

obtain an

equation

of the form

where q, u &#x3E; 0.

Comparison

with

Eq. (1-8),

the

eigenvalue equation

for a2 in Chandrasekhar’s

criterion,

would show that

q(r)

and

t(r)

are the

same

functions

in both

equations,

but that in

general u(r) ~ v(r).

The

same

boundary

conditions of course hold for both

equations.

We will now prove that all the

eigenvalues

l1. of

Eq. (4-2)

are

positive

if and

only

if all the a2’s in

Eq. (1-8)

are

positive.

As

before,

we write

for the lowest

eigenvalues

f1.o and

~o, using q(0)

=

q(R)

=

0,

and

the functional

being

of course the same in both

equations.

Now one

can show

[2]

that there exists an admissible cpo such that

(21)

302 w. J. cocKE

Now suppose 0.

0,

and hence

Therefore 0

implies ao 0,

and one may

similarly

show that

0

implies

0. Hence all the

[.L’S

are

positive

if and

only

if all

the a2’s are

positive,

and hence 82S 0 if and

only

if

cr:

&#x3E; 0.

This shows the

equivalence

of the two

stability

criteria. We now pass

on to consider

applications

of our formalism to certain features of actual models.

5. - APPLICATION TO FEATURES

OF ACTUAL MODELS

Actual

equilibrium

models for a

variety

of

equations

of state have been

fairly thoroughly investigated,

and the

stability

of some of these models has

been tested with Chandrasekhar’s criterion

[19].

Since the results which

we will discuss

[19, 20]

were obtained for

equations

of state of the form

p =

peE)

with no

dependence

on a

co-moving

variable x, we will use p =

p(e) throughout

this section as a

simplifying assumption.

One

aspect

of the results obtained takes the form of curves of total energy

(mass)

of the models as a function of a

parameter

such as the central pres-

sure p~. We write M =

MoCPc).

One may show that

only

one such

function

Mo(pc)

is obtainable for a

given equation

of strate.

A

stability analysis

for a

given pee)

then shows which values of pc are associated with stable

configurations

and which with unstable ones. One feature that

usually

appears is that for a

given equation

of state there is

a

point ~

such that for 0 ~ pc

the

configurations

are

stable,

while

for pc &#x3E;

p~

the

configurations

are unstable. We call the

points p~

« tran-

sition

points )). Further,

the transition

points po

are known to be « mass

peaks

», or maxima of the functions

Mo(pc). However,

it appears that not all mass

peaks

are such transition

points.

We will prove a

plausible

suffi-

cient condition for a mass

peak

to be a

point

of either neutral or unstable

equilibrium.

This will

apply

as well to minima on the mass curves.

Oppen-

heimer and Volkoff

[20]

have

presented

arguments which lead one to the conclusion that all maxima and minima must be such

stability-instability

transition

points.

Calculations

by

Misner and

Zapolsky [19]

have shown

that this is not the case, and we will

explain why Oppenheimer

and Vol-

koff’s

argument

fails.

Références

Documents relatifs

We propose an extension of the principle of virtual work of mechanics to random dynamics of mechanical systems. The total virtual work of the interacting forces and inertial forces

Sellens [21] explained that the unrealistic behavior of the number-based drop-size distribution he obtained with Brzustowski (Equation 23) at small drop diameter

In the random case B → ∞, target probabilities are learned in descending order of their values, and the structure in the spin-configuration space is completely irrelevant:

The experimental accuracy was determined considering the differences between results obtained for data collected at 10 and 90 K, different experimental setups

Top left (A): Prior density distribution from procrystal refined with combined SR- XPD/NPD data without considering thermal displacement; Top right (B): Prior density distribu- tion

This method has recently been applied in [5] for the generation of spectrum-compatible accelerograms as trajectories of a non-Gaussian non- stationary centered random

We have proposed a method to effectively construct the probability density function of a random variable in high dimension and for any support of its probability distribution by

Under regularity conditions, the efficiency properties of the MEM estimator built with the approximate constraint have been studied in [11] and [12], introducing the approximate